RESEARCH ARTICLE

An extensive program of periodic alternative splicing linked to cell cycle progression Daniel Dominguez1,2, Yi-Hsuan Tsai1,3, Robert Weatheritt4, Yang Wang1,2, Benjamin J Blencowe4*, Zefeng Wang1,5*

1Department of Pharmacology, University of North Carolina at Chapel Hill, Chapel Hill, United States; 2Lineberger Comprehensive Cancer Center, University of North Carolina at Chapel Hill, Chapel Hill, United States; 3Program in Bioinformatics and Computational Biology, University of North Carolina at Chapel Hill, Chapel Hill, United States; 4Donnelly Centre and Department of Molecular Genetics, University of Toronto, Toronto, Canada; 5Key Lab of Computational Biology, CAS-MPG Partner Institute for Computational Biology, Chinese Academy of Science, Shanghai, China

Abstract Progression through the mitotic cell cycle requires periodic regulation of function at the levels of transcription, translation, -protein interactions, post-translational modification and degradation. However, the role of alternative splicing (AS) in the temporal control of cell cycle is not well understood. By sequencing the human transcriptome through two continuous cell cycles, we identify ~ 1300 with cell cycle-dependent AS changes. These genes are significantly enriched in functions linked to cell cycle control, yet they do not significantly overlap genes subject to periodic changes in steady-state transcript levels. Many of the periodically spliced genes are controlled by the SR protein kinase CLK1, whose level undergoes cell cycle- dependent fluctuations via an auto-inhibitory circuit. Disruption of CLK1 causes pleiotropic cell cycle defects and loss of proliferation, whereas CLK1 over-expression is associated with various *For correspondence: cancers. These results thus reveal a large program of CLK1-regulated periodic AS intimately [email protected] (BJB); associated with cell cycle control. [email protected] (ZW) DOI: 10.7554/eLife.10288.001 Competing interest: See page 15 Funding: See page 16 Introduction Received: 22 July 2015 Alternative splicing (AS) is a critical step of gene regulation that greatly expands proteomic diversity. Accepted: 24 March 2016 Nearly all (>90%) human genes undergo AS and a substantial fraction of the resulting isoforms are Published: 25 March 2016 thought to have distinct functions (Pan et al., 2008; Wang et al., 2008). AS is tightly controlled, and Reviewing editor: Gene Yeo, its mis-regulation is a common cause of human diseases (Wang and Cooper, 2007). Generally, AS is University of California, San regulated by cis-acting splicing regulatory elements that recruit trans-acting splicing factors to pro- Diego, United States mote or inhibit splicing (Matera and Wang, 2014; Wang and Burge, 2008). Alterations in splicing factor expression have been observed in many cancers and are thought to activate cancer-specific Copyright Dominguez et al. splicing programs that control cell cycle progression, cellular proliferation and migration (David and This article is distributed under Manley, 2010; Oltean and Bates, 2014). Consistent with these findings, several splicing factors the terms of the Creative Commons Attribution License, function as oncogenes or tumor suppressors (Karni et al., 2007; Wang et al., 2014), and cancer- which permits unrestricted use specific splicing alterations often affect genes that function in cell cycle control (Tsai et al., 2015). and redistribution provided that Progression through the mitotic cell cycle requires periodic regulation of gene function that is pri- the original author and source are marily achieved through coordination of protein levels with specific cell cycle stages credited. (Harashima et al., 2013; Vermeulen et al., 2003). This temporal coordination enables timely control

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 1 of 19 Research Article Cell biology Genomics and evolutionary biology

eLife digest Mitosis is a key step in the normal life cycle of a cell, during which one cell divides into two new cells. As a cell progresses through the cell cycle, it must carefully regulate its gene activity to switch particular genes on or off at specific moments. When a gene is activated its sequence is first copied into a temporary molecule called a transcript. These transcripts are then edited to form templates to build . One way that a transcript can be edited is via a process called alternative splicing, in which different pieces of the transcript are cut and pasted together to form different versions of the final template. This allows different instructions to be obtained from a single gene, introducing an added layer of biological complexity. However, the role of alternative splicing in the timing of key events of the cell life cycle is not well understood. Dominguez et al. have now looked for the genes that undergo alternative splicing during the cell cycle. The sequences of gene transcripts produced within human cells were collected while the cells went through two rounds of division. This approach revealed that around 1,300 genes are spliced in different ways at different stages of each cell cycle. Many of these genes were known to play roles in controlling the cell’s life cycle, but few of the genes showed large changes in the amount of total transcript that is generated over time. Dominguez et al. also showed that an enzyme called CLK1 influences about half of the 1,300 periodically spliced genes during the cell cycle. The production of CLK1 is itself carefully controlled throughout the cell cycle, and the enzyme’s activity prevents its own overproduction. Further experiments showed that blocking CLK1’s activity while a cell is replicating its DNA halts the cell cycle, but blocking this enzyme’s activity after the cell had replicated its DNA did not. Given this pivotal role in the cell cycle, Dominguez et al. also examined the role of CLK1 in cancer cells and found that high levels of CLK1 in tumours were linked to lower survival rates. These findings indicate that CLK1 warrants further investigation, particularly in relation to its role in cancer. DOI: 10.7554/eLife.10288.002

of molecular events that ensure accurate chromatin duplication and daughter cell segregation. Peri- odic gene function is conventionally thought to be achieved through stage-dependent gene tran- scription (Bertoli et al., 2013), translation (Grabek et al., 2015), protein-protein interactions (Satyanarayana and Kaldis, 2009), post-translational protein modifications, and ubiquitin-depen- dent protein degradation (Mocciaro and Rape, 2012). Although AS is one of the most widespread mechanisms involved in gene regulation, the relationship between the global coordination of AS and the cell cycle has not been investigated. Major families of splicing factors include the Serine-Arginine rich proteins (SR) proteins and the heterogeneous nuclear ribonucleoproteins (hnRNPs), whose levels and activities vary across cell types. SR proteins generally contain one or two RNA recognition motifs (RRMs) and a domain rich in alternating Arg and Ser residues (RS domain). Generally, RRM domains confer RNA binding specific- ity while the RS domain mediates protein-protein and protein-RNA interactions to affect splicing (Long and Caceres, 2009; Zhou and Fu, 2013). Post-translational modifications of SR proteins, most notably phosphorylation, modulate their splicing regulatory capacity by altering protein locali- zation, stability or activity (Gui et al., 1994; Lai et al., 2003; Prasad et al., 1999; Shin and Manley, 2002). Dynamic changes in SR protein phosphorylation have been detected after DNA damage (Edmond et al., 2011; Leva et al., 2012) and during the cell cycle (Gui et al., 1994; Shin and Man- ley, 2002), suggesting that regulation of AS may have important roles in cell cycle control. However, the functional consequences of SR protein (de)phosphorylation during the cell cycle are largely unclear. Through a global-scale analysis of the human transcriptome at single-nucleotide resolution through two continuous cell cycles, we have identified widespread periodic changes in AS that are coordinated with specific stages of the cell cycle. These periodic AS events belong to a set of genes that is largely separate from the set of genes periodically regulated during the cell cycle at the tran- script level, yet the AS regulated set is significantly enriched in cell cycle- associated functions. We further demonstrate that a significant fraction of the periodic AS events is regulated by the SR pro- tein kinase, CLK1, and that CLK itself is also subject to cell cycle-dependent regulation. Moreover,

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 2 of 19 Research Article Cell biology Genomics and evolutionary biology inhibition or depletion of CLK1 causes pleiotropic defects in mitosis that lead to cell death or G1/S arrest, suggesting that the temporal regulation of splicing by CLK1 is critical for cell cycle progres- sion. The discovery of periodic AS thus reveals a widespread yet previously underappreciated mech- anism for the regulation of gene function during the cell cycle.

Results Alternative splicing is coordinated with different cell cycle phases To systematically investigate the regulation of AS during the cell cycle, we performed an RNA-Seq analysis of synchronously dividing cells using a total of 2.3 billion reads generated across all stages (G1, S, G2 and M) of two complete rounds of the cell cycle (Figure 1—figure supplement 1A). To maximize the detection of regulated AS events, we used the complementary analysis pipelines, MISO and VAST-TOOLS (Katz et al., 2010; Irimia et al., 2014). These pipelines have different detection specificities and employ partially overlapping reference sets of annotated AS events, and therefore afford a more comprehensive analysis when employed together. Both pipelines were used to determine PSI (the percent of transcript with an exon spliced in) and PIR (the percent of tran- scripts with an intron retained). Alternative exons detected by both pipelines had highly correlated PSI values (Figure 1—figure supplement 1G; see below). Consistent with previous results (Bar- Joseph et al., 2008; Whitfield et al., 2002), transcripts from approximately 14.2% (1182) of expressed genes displayed periodic differences in steady-state levels between two or more cell cycle stages (see below). Remarkably, 15.6% (1293) of expressed genes also contained 1747 periodically- regulated AS events, among a total of ~ 40,000 detected splicing events (FDR < 2.5%; Figure 1A and Figure 1—figure supplement 1B,D). Importantly, as has been observed previously for AS regulatory networks (Pan et al., 2004), the majority of genes with periodic AS events did not overlap those with periodic steady-state changes in mRNA expression. This indicates that genes with periodic changes in AS and transcript levels are largely independently regulated during the cell cycle (Figure 1B). Further supporting this conclusion, we did not observe a significant correlation (positive or negative) between exon PSI values and mRNA expression levels for genes with both periodic expression and periodic exon skipping (data not shown). A (GO) analysis reveals that genes with periodic AS, like those with peri- odic transcript level changes (Bar-Joseph et al., 2008; Whitfield et al., 2002), are significantly enriched in cell cycle-related functional categories, including M-phase, nuclear division and DNA metabolic process (Figure 1C; adjusted p<0.05 for all listed categories, FDR<10%) (Supplementary file 1). Similar GO enrichment results were observed when removing the relatively small fraction (10%) of periodically spliced genes that also display significant mRNA expression changes across the cell cycle (Figure 1C). These results thus reveal that numerous genes not previ- ously linked to the cell cycle, as well as previously defined cell cycle-associated genes thought to be constantly expressed across the cell cycle, are in fact subject to periodic regulation at the level of AS (Supplementary file 1 for a full list). Among the different classes of AS analyzed (cassette exons, alternative 5’/3’ splice sites and intron retention [IR]), periodically regulated IR events were over-represented (relative to the back- ground frequency of annotated IR events) by ~ 2.2 fold whereas periodically regulated cassette exons, represent the next most frequent periodic class of AS (p=2.2Â10-16, Fisher’s exact test, Fig- ure 1—figure supplement 1E). Quantitative RT-PCR assays across different cell cycle stages vali- dated periodic IR events detected by RNA-Seq (Figure 1D). Interestingly, one of these IR events is in transcripts encoding aurora kinase B (AURKB), a critical mitotic factor regulated at the levels of transcription, protein localization, phosphorylation and ubiquitination (Carmena et al., 2012; Lens et al., 2010). The AURKB retained intron is predicted to introduce a premature termination codon that elicits mRNA degradation through nonsense mediated decay, and is thus expected to result in reduced levels of AURKB protein. The splicing of the retained intron lags behind changes in the total AURKB mRNA expression (Figure 1E). We computationally corrected levels of fully spliced, protein coding AURKB mRNA by taking into account the fraction of intron-retaining (i.e. non-produc- tive) transcripts across the cell cycle stages (Figure 1E). The expression curve for corrected AURKB mRNA levels is substantially different from total AURKB transcript levels, with a shifted peak coincid- ing with mitosis. Periodically-regulated IR events detected in other genes, including those with

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 3 of 19 Research Article Cell biology Genomics and evolutionary biology

Figure 1. Global detection of periodic cell cycle-dependent alternative splicing. (A) Heat map representation of periodically spliced events. Row- normalized relative PSI values are shown. Diagram below indicates cell cycle phase. (B) Overlap between periodically spliced genes and periodically expressed genes detected by RNA-Seq. (C) Heat map representation of enriched Gene Ontology terms shown as log (p-value). Three gene sets were analyzed separately: all genes with periodic AS, genes with periodic AS only, and genes with both periodic AS and periodic expression. (D) Real-time quantitative PCR analysis of periodic retained introns and total mRNAs for three selected genes. Cells were synchronized by double thymidine block and samples were collected 0, 3, 6, 9, 12 and 15 hr post release. Errors bars represent standard deviation of the mean. Diagram below indicates cell cycle stage. (E) Schematic representation of AURKB AS pattern. Line graph showing the relationship between intron retention and mRNA levels for the AURKB gene across the cell cycle. Percent intron retention (solid red line) across cell cycle was used to determine the fraction of total mRNAs (solid blue line) not containing an intron, i.e. ‘corrected’ mRNA levels (dashed blue line). DOI: 10.7554/eLife.10288.003 The following figure supplement is available for figure 1: Figure supplement 1. Identification of periodic AS by multiple analysis pipelines. DOI: 10.7554/eLife.10288.004

known cell cycle functions such as HMG20B and RAD52, are similarly expected to affect the cell cycle timing of mRNA expression (Figure 1A,D). Collectively, these results provide evidence that the temporal control of retained intron AS provides an important mechanism for establishing the timing of expression of AURKB mRNA and protein, as well as of the timing of expression of additional genes during the cell cycle.

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 4 of 19 Research Article Cell biology Genomics and evolutionary biology

Figure 2. Cell cycle-dependent regulation of CLK1. (A) Immunoblot analysis of proteins involved in splicing regulation in synchronized HeLa cells after release from double thymidine block. (B) Immunoblot analysis of selected proteins in asynchronous HeLa cells or cells arrested at different cell cycle stages. Stably expressed exogenous CLK1 levels were also assessed during the cell cycle (bottom panel). (C) Immunoblot of endogenous CLK1 (top) and exogenously-expressed wild type (CLK1wt) or kinase catalytically inactive (CLK1KD) proteins (bottom) upon treatment with 10 mM TG003. (D) Co- expression of CLK1WT and CLK1KD at different ratios. (E) Immunoprecipitation of CLK1 proteins co-expressed with myc-ubiquitin. Cells were treated with 10 mM TG003 and 10 mM MG132 prior to sample collection. (F) Immunoblot analysis of lysates from cells synchronized upon early S phase (double thymidine) release with or without TG003 treatment. DOI: 10.7554/eLife.10288.005 The following figure supplements are available for figure 2: Figure supplement 1. Periodic expression of RBPs. DOI: 10.7554/eLife.10288.006 Figure supplement 2. Regulation of CLK1 proteins levels during the cell cycle is degradation-dependent. DOI: 10.7554/eLife.10288.007

The SR protein kinase CLK1 fluctuates during the cell cycle Alternative splicing is generally regulated by the concerted action of multiple cis-elements that recruit cognate splicing factors. Consistently, analysis of our RNA-seq data revealed 96 RNA binding proteins (RBPs) with periodic mRNA expression, including RS domain-containing factors like SRSF2, SRSF8, TRA2A and SRSF6 (Figure 2—figure supplement 1A). These 96 RBPs were significantly enriched in the GO term ’splicing regulation’ (adjusted p=10-4, Figure 2—figure supplement 1B), indicating that periodic AS is likely controlled by multiple RBPs. Correlations between these RBPs and periodic splicing events were also identified (Figure 2—figure supplement 1C). For example, SRSF2 expression is significantly correlated with the splicing pattern of a retained intron in the SRSF2 transcript. Further supporting a role for these RBPs in controlling periodic splicing was the identification of RNA motifs bound by a subset of periodically expressed RBPs (Figure 2—figure supplement 1D). To further examine periodic RBP regulation during cell cycle, we measured the abundance of known splicing regulatory proteins at different stages of the cell cycle by immunoblot- ting (Figure 2A). Among the proteins analyzed, CDC-like kinase 1 (CLK1), an important regulator of the Ser/Arg (SR) repeat family of splicing regulators, displayed the strongest cyclic expression

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 5 of 19 Research Article Cell biology Genomics and evolutionary biology peaking at the G2/M phase (Figure 2A,B), consistent with the results of a recent mass-spectrome- try-based screen for cycling proteins (Ly et al., 2014). CLK1 is one of four human CLK paralogs (CLK1-4) and is known to regulate AS via altering the phosphorylation status of multiple SR proteins (Duncan et al., 1997; Jiang et al., 2009; Ninomiya et al., 2011; Prasad et al., 1999). Notably, the levels of other detectable CLK paralogs, as well as members of another SR protein kinase, SRPK1, did not change significantly at the level of RNA and/or protein during the cell cycle (Figure 2B and Supplementary file 1). Given that both CLK1 protein levels and known CLK1 substrates are periodically expressed, we decided to further investigate the role of CLK1 in the context of cell cycle. The levels of total CLK1 mRNA, as well as the levels of specific CLK1 splice variants, did not change significantly during the cell cycle (Figure 2—figure supplement 2A,B) indicating that periodic expression of CLK1 is con- trolled at the level of protein translation and/or turnover. Consistent with this, an exogenously expressed CLK1 protein displayed cell cycle-dependent fluctuations similar to those observed for endogenous CLK1 protein (Figure 2B). Moreover, CLK1 was rapidly degraded upon inhibition of translation by cycloheximide, and this effect was reversed by co-treatment with the proteasome inhibitor MG132 (Figure 2—figure supplement 2C). Additionally, polyubiquitination of Flag-tagged CLK1 was detected following immunoprecipitation with anti-Flag antibody from cells treated with MG132 (Figure 2—figure supplement 2D). These data suggest that the levels of CLK1 protein are controlled by ubiquitin-mediated degradation in a cell cycle-dependent manner. Periodically regulated protein levels are often controlled through negative feedback circuits involving auto-regulatory loops. CLK1 has been reported to auto-phosphorylate on several residues (Ben-David et al., 1991). To investigate whether auto-phosphorylation of CLK1 affects its periodic regulation, we tested whether blocking its kinase activity affects its stability. Inhibition of CLK1 kinase activity using a selective inhibitor, TG003 (Muraki et al., 2004), markedly stabilizes both endogenous and exogenously expressed CLK1 proteins (Figure 2C). Moreover, activity-dependent destabilization of CLK1 was observed with a wild type (WT) protein, but not with a catalytically inac- tive (KD) mutant (Figure 2C, left panel). We further observe that WT CLK1 is rapidly degraded upon cycloheximide treatment, whereas the KD mutant is more stable (Figure 2—figure supplement 2C). We next tested whether CLK1 activity is sufficient to trigger its own degradation by co-expressing KD CLK1 with increasing amounts of WT CLK1. As expected, increasing amounts of WT CLK1 reduces levels of KD CLK1 (Figure 2D). Consistent with these results, WT CLK1 is more highly polyu- biquitinated compared to the KD mutant (Figure 2E, compare lanes 2 to 4), and treatment with TG003 reduces polyubiquitination levels (Figure 2E, lane 2 vs. 3 and lane 4 vs. 5). Decreased polyu- biquitination of WT CLK1 is more prevalent than is apparent upon TG003 treatment, as CLK1 is sta- bilized by TG003 inhibition and thus more total Flag-CLK1 is immunoprecipitated (Figure 2E). To further examine whether this auto-feedback loop is required for changes in CLK1 protein levels dur- ing the cell cycle, we treated synchronized cells with TG003 (or DMSO as a control) and measured CLK1 protein levels. We observed that CLK1 inhibition prevents its turnover after the G2/M phase for both endogenous and exogenously expressed kinases (Figure 2F and Figure 2—figure supple- ment 2E). Taken together, these results provide strong evidence that CLK1 protein levels are con- trolled by ubiquitin-mediated proteolysis in a cell cycle stage-specific manner, and that an activity- dependent negative feedback loop is required for this periodic regulation. These results further sug- gest that changes in the levels of CLK1 could account for many of the periodically regulated AS tran- sitions we have detected during the cell cycle.

CLK1 regulates AS events in genes with critical roles in cell cycle control RNA-Seq analysis of cells treated with TG003 revealed 892 AS events (in 665 genes) that significantly change after CLK1 inhibition (Figure 3A), including known CLK1-regulated splicing events (e.g. exon 4 of CLK1 [Duncan et al., 1997]). It is worth noting that TG003 can also inhibit CLK4 (although to a lesser extent than inhibition of CLK1). However, RNAi of CLK1 is sufficient to recapitulate the pheno- type of CLK1/4 inhibition (see below and Figure 4)(Fedorov et al., 2011; Muraki et al., 2004). Intron retention and cassette exons are the most overrepresented types of AS affected by TG003 (Figure 3A). Most (70%) of the CLK1-regulated exons display increased skipping upon CLK1 inhibi- tion, whereas 87% of CLK1-regulated introns show increased retention (Figure 3—figure supple- ment 1B and C), consistent with a recently reported role for CLK1 in the regulation of retained

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 6 of 19 Research Article Cell biology Genomics and evolutionary biology

Figure 3. CLK1 regulates a network of genes that control cell cycle progression. (A) Identification of endogenous CLK1 targets by RNA-Seq. Numbers of different AS types affected by treatment with the CLK1 inhibitor TG003 (left graph). SE, skipped exon; RI, retained intron; A3E, alternative 3’ exon; A5E, alternative 5’ exon. Fraction of total analyzed events that were affected by TG003 treatment (right graph). (B) Validation of TG003-responsive AS events by semi-quantitative RT-PCR. The bar graph shows the max-delta PSI for each AS event tested in a 24-hr time course of inhibition with 20mM TG003. (C) Representation of cell cycle control genes with CLK1-dependent AS events, organized by cell cycle phase and function. (D) Schematic representation of CHEK2 alternative splicing, showing that exon 9 encodes a region overlapping the kinase domain (upper panel). Semi-quantitative RT-PCR assessment of CHEK2 isoforms after treatment with TG003 or over-expression of the indicated factors (lower left panel). RNAi of SRSF1 in cells and subsequent analysis of CHEK2 splicing by semi-quantitative RT-PCR (lower right panel). PSI values are shown below gel. (E) Normalized CENPE total mRNA expression during an unperturbed cell cycle (triangle denotes mitosis, left panel) and diagram of CENPE splicing (right). TG003-treatment Figure 3 continued on next page

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 7 of 19 Research Article Cell biology Genomics and evolutionary biology

Figure 3 continued of HeLa cells released from G1/S arrest followed by semi-quantitative RT-PCR analysis of CENPE isoforms (bar graph). (F) Schematic of RNA-Seq analysis of CLK1 inhibition during cell cycle (left). AS events that were identified as being differentially regulated between G1 and G2 phase (top bar of bar graph) and number of events that were blocked by the indicated conditions (bottom 4 bars). DOI: 10.7554/eLife.10288.008 The following figure supplement is available for figure 3: Figure supplement 1. CLK1 regulates a network of genes that control cell cycle progression. DOI: 10.7554/eLife.10288.009

introns (Boutz et al., 2015). Of nine analyzed TG003-affected AS events detected by RNA-Seq anal- ysis, all were validated by semi-quantitative RT-PCR assays (Figure 3B). These observations indicate that CLK1 inhibition mainly suppresses splicing, consistent with a general requirement for phosphor- ylation of SR proteins to promote splicing activity (Irimia et al., 2014; Prasad et al., 1999; Tsai et al., 2015). Importantly, there is a significant overlap between genes with cell cycle periodic AS events (Figure 1) and those with CLK1-regulated AS events, involving 156 genes (p=8.5Â10-10, hyper-geometric test). In contrast, consistent with the results in Figure 1, we do not observe a signif- icant overlap between genes containing CLK1-regulated AS events and periodically expressed genes. These results thus support a widespread and rapidly acting role for CLK1 in controlling cell cycle-regulated AS. Indeed, CLK1 inhibition induces rapid (within 3–6 hr) changes in AS among sev- eral analyzed cases (Figure 3—figure supplement 1D). Supporting an important role for CLK1 in cell cycle progression, genes whose AS levels are affected by CLK1 inhibition are significantly enriched in the GO terms cell cycle phase, M-phase, DNA metabolic processes, nuclear division, DNA damage response, and cytokinesis (adjusted p<0.05 and FDR <20% for all listed GO terms; full list in Supplementary file 2). The affected genes function at various stages of cell cycle including the G1/S transition (Figure 3C). Mitotic processes were, however, associated with the largest number of CLK1-target genes with AS changes and included examples that function in duplication (CEP70, CEP120, CEP290, CEP68, CDK5RAP2), metaphase and anaphase (e.g. CENPK, CENPE, CENPN), and cytokinesis (e.g. SEPT2, SEPT10, ANLN) (additional examples in Figure 3C). To further investigate the functional consequence of CLK1-dependent AS, we selected two exam- ples in genes that have important roles in the cell cycle: checkpoint kinase 2 (CHEK2), a tumor sup- pressor that controls the cellular response to DNA damage and cell cycle entry (Paronetto et al., 2011; Staalesen et al., 2004), and centromere-associated protein E (CENPE), a kinetochore-associ- ated motor protein that functions in alignment and segregation during mitosis (Kim et al., 2008). We detected a TG003 dose-dependent increase in CHEK2 exon 9 inclusion, whereas overexpression of WT CLK1 induced exon 9 skipping, an event that removes the CHEK2 kinase domain (Figure 3D). Expression of WT CLK1 in the presence of TG003, or a catalytically inac- tive CLK1, had little to no effect on the splicing of this exon (Figure 3D, bottom panels), indicating that the catalytic activity of CLK1 is essential for regulating CHEK2 AS. Over-expression of the SR protein splicing regulator SRSF1, a known target of CLK1 (Prasad et al., 1999), had a similar effect as over-expression of CLK1, resulting in CHEK2 exon 9 skipping, whereas knockdown of SRSF1 had the opposite effect (Figure 3D, right panel). Furthermore, the activation of CHEK2 requires homodi- merization (Shen et al., 2004), and we observe that the CHEK2 isoform lacking exon 9 still interacts with full-length protein (Figure 3—figure supplement 1E), suggesting that this CLK1-regulated iso- form may function in a dominant-negative manner to attenuate CHEK2 activity. CENPE is known to be tightly controlled at multiple levels (including transcription, localization, phosphorylation and degradation), and disruption of its regulation leads to pronounced mitotic defects. CENPE AS generates long and short isoforms (Supplementary file 2), with the predominant variant being the short isoform that lacks amino acids 1972–2068. Inhibition of CLK1 rapidly shifts CENPE splicing to produce predominantly the long isoform (Figure 3E), and this is accompanied by a reduction in CENPE protein levels during G2/M phase, presumably due to the instability of the long isoform (Figure 3E). These data thus show that CLK1 controls the AS of major cell cycle regula- tors, and therefore suggest that inhibition of CLK1 may alter cell cycle progression. To investigate this, we next performed an RNA-Seq analysis of synchronized cells at G1 and early G2 (when CLK1 accumulates), following inhibition of CLK1 with TG003. As a specificity control, we

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 8 of 19 Research Article Cell biology Genomics and evolutionary biology

Figure 4. CLK1 is required for cell cycle progression and proliferation. (A) Immunoblot analysis of CLK1 proteins after stable shRNA knockdown in HeLa cells. Bottom, DNA content as measured by propidium iodide staining following flow cytometry. (B) Immunofluorescence microscopy of A549 cells depleted of CLK1 by shRNA (top row), cells treated with 10 mM TG003 for 12 hr (middle row), and a control treatment with DMSO (bottom row); green: tubulin, red: emerin (nuclear envelope), and blue: DAPI. Scale bar 10 mm. Right bar graph shows the quantification of multinucleated cells. p values determined using Student’s t-test. (C) Static frames from a live-cell high-content imaging movie of HeLa cells expressing Histone H2B-GFP and treated with TG003 (top panel). Time after start of Figure 4 continued on next page

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 9 of 19 Research Article Cell biology Genomics and evolutionary biology

Figure 4 continued the experiment is indicated; EP, end point ( ~ 960 min). TG003 treated cells with apparent cell division defects (indicated by arrowheads in the bottom field) are shown in two independent fields. (D) Synchronized HeLa cells were treated with 20 mM TG003 at the indicated time points (0, 5, and 10 hr) and analyzed by propidium iodide staining and flow cytometry to measure DNA content. Percent of 2N (lower bar graph) and 4N (upper bar graph) cells were quantified at each time point as indicated in the treatment scheme (top). (E) Colony formation assay of HeLa cells depleted of CLK1 by shRNA, or continuously treated with TG003 or KHCB-19 at the indicated concentrations. (F) Box plot representation of CLK1 mRNA expression levels in paired normal and tumorous kidney tissue. 72 cases were analyzed. (G) Kaplan-Meier plot showing survival differences between patients with kidney tumors with high CLK1 (red, upper quartile) or reduced CLK1 (blue, lower three quartiles) expression. (H) Number of cancer-associated AS events that are also regulated by CLK1 in different tumor types. BRCA, Breast invasive carcinoma; COAD, Colorectal adenocarcinoma; KIRC, Kidney renal clear cell carcinoma; LUAD, Lung adenocarcinoma; LUSC, Lung squamous cell carcinoma; LIHC, liver hepatocellular carcinoma. DOI: 10.7554/eLife.10288.010 The following figure supplements are available for figure 4: Figure supplement 1. Loss of CLK1 results in cell cycle defects in multiple cell types. DOI: 10.7554/eLife.10288.011 Figure supplement 2. CLK1 mis-regulation in human cancer. DOI: 10.7554/eLife.10288.012

performed a parallel RNA-Seq analysis using a structurally distinct CLK1 inhibitor, KHCB-19 (Fedorov et al., 2011). Strikingly, of 1498 AS events that change between G1 (t=0) and G2 (t=6) ~ 94% display reduced changes following treatment with the two drugs (Figure 3F), with 65% com- monly affected by both drugs, thus supporting an important role for CLK1 in controlling cell-cycle dependent splicing.

CLK1 is required for normal mitosis and cell proliferation Given that CLK1 regulates the AS of many cell cycle factors (Figure 3C), we next examined whether it is necessary for cell cycle progression. Knockdown of CLK1 using shRNAs led to an accumulation of cells with 4N DNA content in multiple cell types, specifically HeLa, H157, and A549 (Figure 4A, Figure 4—figure supplement 1A,B), and the extent of this accumulation correlated with the degree of knockdown (Figure 4—figure supplement 1A). We also observed a significant increase in multi- nucleation, a common consequence of defective chromosome segregation or cytokinesis, following shRNA-knockdown or TG003 inhibition of CLK1 in the treated cells (Figure 4B and Figure 4—figure supplement 1C). To visualize the effect of CLK1 inhibition on mitosis at a single cell level, we per- formed time-lapse high-content microscopy on live cells stably expressing a GFP-histone 2B fusion protein, to track changes in chromatin. TG003-treated cells entered mitosis normally, as measured

Video 2. Live-cell imaging of control HeLa cells stably Video 1. Live-cell imaging of control HeLa cells stably expressing GFP-H2B. Cells were synchronized by expressing a GFP-H2B. Cells were synchronized by single-thymidine block and released into 1 mM of single-thymidine block and released and imaged at TG003 and imaged at 10X magnification (every ~ 15 10X magnification (every ~ 15 min) for 960 min. min) for 960 min. DOI: 10.7554/eLife.10288.013 DOI: 10.7554/eLife.10288.014

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 10 of 19 Research Article Cell biology Genomics and evolutionary biology

Video 3. Zoom of live-cell imaging of control HeLa Video 4. Zoom of live-cell imaging of TG003-treated cells. Data associated with Figure 4D. HeLa cells. Data associated with Figure 4. DOI: 10.7554/eLife.10288.015 DOI: 10.7554/eLife.10288.016

by nuclear envelope breakdown, but displayed delayed or aberrant cytokinesis, typically resulting in multi-polar divisions, increased time in metaphase, failure to undergo chromatin de-condensation and eventual cell death (Figure 4C and Videos 1–5). To further determine at what cell cycle stage CLK1 activity is required, we inhibited CLK1 using TG003 at different time points after early S phase release. Consistent with the imaging data, both control and TG003-treated cells entered mitosis nor- mally, as measured by 4N DNA content. However, inhibition of CLK1 before late S-phase impaired progression through mitosis, whereas cells treated 5 hr after early S phase release underwent a round of normal mitotic division, although failed to enter the next cell cycle (Figure 4D and Fig- ure 4—figure supplement 1D). These results suggest that the primary defects caused by CLK1 inhi- bition occur in late S-phase and G2 phase, which is when CLK1 levels normally begin to rise (Figure 2A). This conclusion is further supported by the observation that CLK1-dependent AS tar- gets, such as those detected in HMMR and CENPE, are periodically expressed during cell cycle and peak during G2 and M phase (Figure 4—figure supplement 1E). Taken together with the earlier results, these data support an important and multifaceted role for CLK1 in the control of cell cycle progression through its function in the global regulation of periodic AS.

CLK1 expression and CLK1-regulated AS is altered in kidney cancer The importance of CLK1 for faithful progression through the cell cycle suggests it may play a role the control of cell proliferation in cancer. Supporting this, shRNA knockdown or chemical inhibition of CLK1 with TG003 or KHCB-19 in HeLa cells results in a near complete block in cell proliferation, as measured by anchorage- dependent and -independent colony formation assays (Figure 4E and Figure 4—figure sup- plement 1F). As mentioned above, CLK1 is likely the primary target of inhibition in these experiments since RNAi of CLK1 recapitulates the phenotype seen with these chemical inhibitors. Using RNA-Seq data from the Cancer Genome Atlas (TCGA) (Cancer Genome Atlas Network, 2013) we observe that CLK1 displays significantly higher expression in 72 kidney tumors compared to matched normal tissue samples (Figure 4F, p=10-5, Kolmogorov-Smir- nov test). Consistently, most CLK1-controlled AS Video 5. Zoom of live-cell imaging of TG003-treated events that are altered in tumors have expected HeLa cells. Data associated with Figure 4. splicing changes (Figure 4—figure supplement DOI: 10.7554/eLife.10288.017 2A). Furthermore, patients with tumors that

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 11 of 19 Research Article Cell biology Genomics and evolutionary biology have elevated CLK1 expression (i.e. the upper quartile of all samples) have significantly lower survival rates relative to other patients in the comparison group (Figure 4G, p=0.007). While there was also an increase in CLK2, CLK3 and CLK4 mRNA expression in these tumors, CLK1 displayed the highest relative mRNA expression levels compared to CLK2 and CLK3 (Figure 4—figure supplement 2B). Levels of the CLK4, which is ~ 80% identical to CLK1, did not correlate with survival differences despite its increased levels in tumors (Figure 4—figure supplement 2C). Consistent with this, CLK1- regulated AS events, as defined by the RNA-Seq analysis in Figure 3, were also altered across multi- ple tumor types, including breast, colon, lung and liver (Figure 4H and Figure 4—figure supple- ment 2D). These data are further consistent with a multi-faceted role for CLK1 in regulating cell cycle progression, and also suggest that CLK1 contributes to increased cell proliferation in cancer, at least in part through its role in controlling periodic AS.

Discussion Previous studies have shown that splicing and the cell cycle are intimately connected processes. Indeed, cell cycle division (CDC) loci originally defined in S. cerevisiae, namely cdc5 and cdc40, were subsequently shown to encode spliceosomal components (Ben-Yehuda et al., 2000; McDonald et al., 1999). Moreover, genome-wide RNAi screens for new AS regulators of apoptosis genes in human cells revealed that factors involved in cell-cycle control, in addition to RNA process- ing components, were among the most significantly enriched hits (Moore et al., 2010; Tejedor et al., 2015). An RNAi screen performed in Drosophila cells for genes required for cell-cycle progression identified numerous splicing components (Bjo¨rklund et al., 2006) as well as a Drosoph- ila ortholog of CLK kinases, Darkener of apricot Doa (Bettencourt-Dias et al., 2004). In other stud- ies, negative control of splicing during M phase was shown to be dependent on dephosphorylation of the SR family protein, SRSF10 (Shin and Manley, 2002), and the mitotic regulator aurora kinase A (AURKA) was shown to control the AS regulatory activity of SRSF1 (Moore et al., 2010). Our study shows for the first time that AS patterns are subject to extensive periodic regulation, in part via a global control mechanism involving cell cycle fluctuations of the SR protein kinase CLK1. At least one likely function of this periodic AS regulation is to control the timing of activation of AURKB (Fig- ure 1), as well as of numerous other key cell cycle factors shown here to be subject to periodic AS. The definition of an extensive, periodically-regulated AS program in the present study thus opens the door to understanding the functions of an additional layer of regulation associated with cell cycle control and cancer. Since many RBPs are found to be periodically expressed (Figure 4A), it is likely that other aspects of mRNA metabolism are also coordinated with cell cycle stages. For example, differential degrada- tion could potentially contribute to the observed periodic fluctuation of splice isoforms during cell cycle. The degradation of mRNA is closely linked with alternative polyadenylation, which has emerged as a critical mechanism that controls mRNA translation and stability. Generally, shortened 3’ UTRs are found in rapidly dividing cells and more aggressive cancers (Mayr and Bartel, 2009). This study identified 94 cases periodic alternative poly-A site usage (data not shown) in genes known to regulate cell cycle and/or proliferation, including SON, CENPF and EPCAM (Ahn et al., 2011; Bomont et al., 2005; Chaves-Perez et al., 2013). In addition, many mRNAs have recently been found to be translated in a cell cycle dependent fashion (Aviner et al., 2015; Maslon et al., 2014; Stumpf et al., 2013). Interestingly, a fraction of periodically translated genes are also periodically spliced, including key regulators of cell cycle (e.g., AURKA, AURKB, TTBK1 and DICER1). This obser- vation is consistent with recent findings that the regulation of AS and translation may be coupled (Sterne-Weiler et al., 2013). In summary, we have demonstrated that AS is subject to extensive tem- poral regulation during the cell cycle in a manner that appears to be highly integrated with orthogo- nal layers of cell cycle control. These results thus provide a new perspective on cell cycle regulation that should be taken into consideration when studying this fundamental biological process, both in the context of normal physiology and diseases including cancers.

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 12 of 19 Research Article Cell biology Genomics and evolutionary biology Materials and methods Cell culture and synchronizations HeLa (a kind gift from J. Trejo), HEK 293T (from ATCC CRL-3216) and A549 (kind gift from W. Kim) cells were maintained in DMEM (Gibco) medium supplemented with 10% FBS (Gibco). All cells were

cultured in humidified incubators with 5% CO2. Cell cycle synchronization was adapted from the pro- tocol of Whitfield et al. (Whitfield et al., 2002); ~ 750,000 log phase HeLa cells were plated in 15 cm dishes in complete media and allowed to attach for 16 hr, reaching <30% confluence. Cells were subsequently treated with 2 mM thymidine (Sigma-Aldrich, St. Louis, MO) for a total of 18 hr, washed 2 times with 1xPBS, and supplemented with fresh complete media for 10 hr. 2 mM thymi- dine was subsequently added for a second block of 18 hr and washed as described previously. Mitotic block was performed by double thymidine arrest (as above) and release in fresh media for 3.5 hr followed by addition of nocodazole 100 mM (Sigma) for 10 hr. G1 block was performed by serum starvation for 72 hr in DMEM containing 0.05% FBS. For RNA-Seq, cells (both adherent and detached) were harvested every 1.5 hr for 30 hr and frozen immediately for purification of total RNA. To block the activity of CLK1, cells were treated with TG003 (Sigma), KHCB-19 (Tocris, Bristol, UK). To block activity of the proteasome cells were treated with MG132 (Sigma). Drugs were-suspended in DMSO and added to growing cultures at the indicated concentrations and times.

Flow cytometry and cell cycle analysis Cells were harvested with trypsin treatment, washed 2 times in cold 1xPBS and subsequently fixed in 80% ice cold ethanol for at least 4 hr. Cells were then washed twice with 1xPBS and suspended in propidium iodide/RNase staining buffer (BD Pharmingen, cat # 550825). Cells were analyzed by flow cytometry to count 10,000 cells that satisfied gating criteria. Data collected were analyzed using ModFit software to discern 2N (G1), S-phase, and 4N (G2 and M) composition.

Mapping and filtering of RNA-Seq data RNA-Seq reads were mapped to the (build hg19) using the MapSplice informatics tool with default parameters (Wang et al., 2010). The mapped reads were further analyzed with Cuf- flinks to calculate the level of with FPKM (Fragments Per Kilobase per Million mapped reads) (Trapnell et al., 2010). The levels of alternatively spliced isoforms were quantified with MISO (Mixture-of-Isoforms) probabilistic framework (Katz et al., 2010) using the annotated AS events for human hg19version 2. The levels of alternatively spliced isoforms were also quantified with VAST-TOOLS using the event annotation as previously described (Irimia et al., 2014). Each AS event was assigned a PSI or PIR value to represent the percent of transcripts with the exon spliced in, or the intron retained, respectively.

Identification of periodic AS For identification of periodic AS raw PSI/PIR values were normalized as: F s À F s normalized s n min ÀÁF n ¼ s F max

where s = 1 to 32,109 for all splicing events; n = 1 to 14 for the 14 samples; F min is the minimum and F max is the maximum PSI value among the 14 samples. To identify periodic AS events, normalized gene expression values (normalized FPKM values as en) for the well-known periodic gene, CCNB1, CCNA2, CCNB2, and CENPE, were used as a starting point to subsequently add curves with broader or sharper peaks as well as shifted to left or right, resulting in 7 periodic expression curves that cover all the phases of cell cycle (Figure 1—figure sup- plement 1A). We term these ’ideal seed curves’, which capture intermittent peak times and phase shifts that were not well represented within the initial known periodic genes. To identify genes with similar splicing patterns across the cell cycle, we computed the EuclideanDistanceED of each AS event s to the model seed curves m as follows:

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 13 of 19 Research Article Cell biology Genomics and evolutionary biology

14 ED normalized em normalized s m;s ¼ Xj ð n ÞÀ ð F nÞj n¼1 where m = 1 to 7 for all model seed curves, s = 1 to 32,109 for all AS events. Based on the ranking of distance, a similar cutoff of ED  2:75 was set as a minimum requirement for periodic AS. Lastly, we calculated a false discovery rate (FDR) by shuffling PSI values across the 14 time points 10,000 times and calculating how often a random shuffle had a better periodic score than the true periodic score for that event. A maximum FDR of 2.5% was required for a splicing event to be periodic.

Heat maps, correlations, GO-term analysis, overlap analysis and statistics Heat maps, hierarchical clustering, and Pearson correlations were generated using GENE-E (www. broadinstitute.org/cancer/software/GENE-E/). All heat maps shown are row-normalized for presenta- tion purposes. Spearman’s rank correlation with average linkage was used for clustering. DAVID (http://david.abcc.ncifcrf.gov/gene2gene.jsp) was used for all gene ontology enrichment; terms shown are for biological process (GOTERM_BP_FAT). To test for significance in overlap analysis, overlapping genes in two data sets (i.e. TG003-treatment and periodic AS) and a background set of only co-detected events was used (i.e. genes detected in both experiments). Significance of overlap- ping gene sets was assessed using the hyper-geometric test. For overlap and correlation analysis between VAST-TOOLS and MISO we used two MISO AS event annotations (HG18 and HG19), due to differences in the input annotation files for these two pipelines. To identify the 4,343 overlapping exons in the heatmap, we used MISO hg19v1 annotations and MISO hg18 annotations, and the VAST-TOOLS annotations. Student’s t-test was used to measure significant in cell cycle defects (multi-nucleation and flow-cytometry) as well as semi-quantitative RT-PCR assessment of splice var- iants. To identify AS events blocked by TG003 or KHCB19, a Student’s t-test statistic was used. If a change between G1/S (t=0) and G2 (t=6) was significant, but not significant in the presence of inhibi- tors we consider that event to be blocked. For over-representation of periodic introns, we per- formed Fisher’s exact test in a 2x2 contingency table as compared to skipped exons. For differences in expression of CLK1 mRNAs in kidney cancers a Kolmogorov-Smirnov test was performed. For sur- vival differences, the survdiff function in the R survival package was used (as discussed in methods below).

Plasmid construction, transfections and RNAi The expression constructs were generated by cloning the cDNA of CLK1 into pCDNA3 (for transient expression) or pCDH (for stable transfection) backbones with different epitope tags (HA or Flag) at the N- or C-terminus. The Myc-His-Ubiquitin expression vector is a gift from Dr. Gary Johnson’s lab, and the Histone H2B-GFP expression vector is gift from Dr. Angelique Whitehurst’s lab. Plasmid transfections were performed using Lipofectamine 2000 (Invitrogen) according to the manufacturer’s protocol. The lentiviral vectors of shRNAs were obtained from Addgene in pLKO.1 TRC cloning plas- mid through UNC core facility as part of mammalian gene knockdown consortium. Lentiviral infec- tions were performed according to the manufacturer’s instruction from System Biosciences (SBI).

Semi-quantitative RT-PCR assays for monitoring splicing Cells transfected with shRNA constructs or treated with TG003 were lysed and total RNA was extracted using the Trizol method (Life Technologies). Purified RNA was treated with 1U of RNAse- free DNAase (Promega) for 1 hr at 37˚C and reverse transcribed using random hexamer cDNA prep- aration kit (Applied Biosystem). One-tenth of the RT product was used as the template for PCR amplification (25 cycles of amplification, with a trace amount of Cy5-dCTP in addition to non-fluores- cent dNTPs) using gene specific primers listed in Supplementary file 4. The resulting gels were scanned with a Typhoon 8600 Imager (GE Healthcare), and analyzed with ImageQuant 5.2 software (Molecular Dynamics/GE Healthcare). Real time PCR was carried out using the SYBR Green kit (Invi- trogen) and GAPDH as an internal control.

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 14 of 19 Research Article Cell biology Genomics and evolutionary biology Immunoblotting and immunoprecipitation

Proteins were extracted in lysis buffer (CHAPS 1% w/v, 150 mM NaCl, 50 mM MgCl2 with protease inhibitor), resolved by SDS-PAGE and transferred onto PVDF membrane. For immunoprecipitaion experiments to detect ubiquitination, cells were co-transfected with Flag-CLK1 and myc-ubiquitin constructs as above. 36 hr later, TG003 (20 mM) was added for 18 hr. 4 hr prior to harvesting, 10 mM of MG132 was added to the media. Proteins were extracted in lysis buffer as above with the addition of NEM. Incubation with EZ-View FLAG Beads (Sigma) was performed for 2 hr at 4˚C. Samples were extensively washed according to the manufacturer’s protocol and subjected to immunoblotting. Antibodies and dilutions are listed in Supplementary file 4.

Immunofluorescence and high-content live-cell imaging For immunofluorescence microscopy, cells were plated on glass coverslips coated with poly-L-Lysine. Cells were then washed twice with 1xPBS, fixed with 4% formaldehyde (Sigma), permeabilized with 0.05% Triton X-100 (Promega) and blocked with 3% BSA (Fisher); all dilutions were made in 1XPBS. For live cell imaging, HeLa cells transduced with Histone H2B-GFP was stably selected as described previously (Cappell et al., 2010). Cells were plated in a 6-well format and treated with 2 mM thymi- dine for 24 hr, subsequently washed and released in fresh complete medium with or without TG003 (20 mM). Cells were imaged using the BD Pathway Microscope with a 10X objective.

Colony formation assays HeLa cells stably producing shRNAs targeting CLK1 or control shRNAs were plated at low density (1000 cells/6 cm2) in standard culture medium and allowed to proliferate for 9 days. Cells were then fixed and stained with crystal violet at room temperature. The dried plates were used for estimations of colony diameter and number.

Kidney cancer analysis of CLK1 and alternative splicing RNA-Seq data from the The Cancer Genome Atlas (Ciriello et al., 2013) was processed as previ- ously described (Tsai et al., 2015). Briefly, for mRNA expression, RSEM expression values for the indicated genes were analyzed in 79 paired KIRC (tumor and normal) samples and the paired ks-test was used to test significance. For alternative splicing analysis data from BRCA: Breast invasive carci- noma, COAD: Colorectal adenocarcinoma, KIRC: Kidney renal clear cell carcinoma, LUAD: Lung ade- nocarcinoma, LUSC: Lung squamous cell carcinoma, LIHC: liver hepatocellular carcinoma were analyzed through the MISO pipeline as described above (Tsai et al., 2015). Relapse-free survival was analyzed using Kaplan Meier plots. All plots and statistical analyses (survdiff) were generated using the R package version 3.1.1 survival function.

Acknowledgements We thank Dr. Jan Prins and Darshan Singh for helpful suggestions in analyzing RNA-Seq data, and Drs. Angelique Whitehurst, William Marzluff and Jean Cook for sharing reagents. We thank Drs. Jun Zhu and Ting Ni for advice and assistance in generating strand-specific RNA library. Dr. Rebecca Sin- nott provided assistance in live cell imaging experiments. We thank Drs. William Marzluff, Chris Burge, Jean Cook, Michael Emanuele and Mauro Calabrese for helpful comments on the manuscript. This work is supported by an NIH grant (R01CA158283) and Jefferson-Pilot fellowship to ZW, and by grants from the Canadian Institutes of Health Research to BJB. BJB holds the Banbury Chair of Medi- cal Research at the University of Toronto.

Additional information

Competing interests BJB: Reviewing editor, eLife.. The other authors declare that no competing interests exist.

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 15 of 19 Research Article Cell biology Genomics and evolutionary biology Funding Funder Grant reference number Author National Cancer Institute R01CA158283 Zefeng Wang

The funders had no role in study design, data collection and interpretation, or the decision to submit the work for publication.

Author contributions DD, Conceived the project design, Performed the experiments, Analyzed data, Wrote the paper; Y- HT, YW, Analyzed data, Acquisition of data; RW, Analyzed data, Drafting or revising the article; BJB, Analyzed data, Wrote the paper; ZW, Conceived the project design, Analyzed data, Wrote the paper

Author ORCIDs Zefeng Wang, http://orcid.org/0000-0002-6605-3637

Additional files Supplementary files . Supplementary file 1. Periodic splicing events identified during the HeLa cell cycle and associated GO terms. Data associated with Figure 1. DOI: 10.7554/eLife.10288.018 . Supplementary file 2. AS events altered by CLK1 inhibition (TG003 10 mM for 16 hr) in 293T cells and associated GO terms. Data associated with Figure 3. DOI: 10.7554/eLife.10288.019 . Supplementary file 3. AS events altered by CLK1 inhibition (TG003 10 mM, KHCB19 10 mM) in syn- chronized HeLa cells. Data associated with Figure 3. DOI: 10.7554/eLife.10288.020 . Supplementary file 4. List of PCR primers, shRNA sequences and antibodies used in this study. DOI: 10.7554/eLife.10288.021

Major datasets The following dataset was generated:

Database, license, and accessibility Author(s) Year Dataset title Dataset URL information Dominguez D, Tsai 2016 An extensive program of periodic http://www.ncbi.nlm.nih. Publicly available at Y-H, Wang Z alternative splicing linked to cell gov/geo/query/acc.cgi? the Gene Expression cycle progression acc=GSE81485 Omnibus NCBI website (accession no: GSE81485)

References Ahn EY, DeKelver RC, Lo MC, Nguyen TA, Matsuura S, Boyapati A, Pandit S, Fu XD, Zhang DE. 2011. SON controls cell-cycle progression by coordinated regulation of RNA splicing. Molecular Cell 42:185–198. doi: 10. 1016/j.molcel.2011.03.014 Aviner R, Shenoy A, Elroy-Stein O, Geiger T. 2015. Uncovering hidden layers of cell cycle regulation through integrative multi-omic analysis. PLoS Genetics 11:e1005554. doi: 10.1371/journal.pgen.1005554 Bar-Joseph Z, Siegfried Z, Brandeis M, Brors B, Lu Y, Eils R, Dynlacht BD, Simon I. 2008. Genome-wide transcriptional analysis of the human cell cycle identifies genes differentially regulated in normal and cancer cells. Proceedings of the National Academy of Sciences of the United States of America 105:955–960. doi: 10. 1073/pnas.0704723105 Ben-David Y, Letwin K, Tannock L, Bernstein A, Pawson T. 1991. A mammalian protein kinase with potential for serine/threonine and tyrosine phosphorylation is related to cell cycle regulators. The EMBO Journal 10:317– 325. Ben-Yehuda S, Dix I, Russell CS, McGarvey M, Beggs JD, Kupiec M. 2000. Genetic and physical interactions between factors involved in both cell cycle progression and pre-mRNA splicing in Saccharomyces cerevisiae. Genetics 156:1503–1517.

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 16 of 19 Research Article Cell biology Genomics and evolutionary biology

Bertoli C, Skotheim JM, de Bruin RA. 2013. Control of cell cycle transcription during G1 and S phases. Nature Reviews. Molecular Cell Biology 14:518–528. doi: 10.1038/nrm3629 Bettencourt-Dias M, Giet R, Sinka R, Mazumdar A, Lock WG, Balloux F, Zafiropoulos PJ, Yamaguchi S, Winter S, Carthew RW, Cooper M, Jones D, Frenz L, Glover DM. 2004. Genome-wide survey of protein kinases required for cell cycle progression. Nature 432:980–987. doi: 10.1038/nature03160 Bjo¨ rklund M, Taipale M, Varjosalo M, Saharinen J, Lahdenpera¨ J, Taipale J. 2006. Identification of pathways regulating cell size and cell-cycle progression by RNAi. Nature 439:1009–1013. doi: 10.1038/nature04469 Bomont P, Maddox P, Shah JV, Desai AB, Cleveland DW. 2005. Unstable microtubule capture at kinetochores depleted of the centromere-associated protein CENP-F. The EMBO Journal 24:3927–3939. doi: 10.1038/sj. emboj.7600848 Boutz PL, Bhutkar A, Sharp PA. 2015. Detained introns are a novel, widespread class of post-transcriptionally spliced introns. Genes & Development 29:63–80. doi: 10.1101/gad.247361.114 Cancer Genome Atlas Research Network. 2013. Comprehensive molecular characterization of clear cell renal cell carcinoma. Nature 499:43–49. doi: 10.1038/nature12222 Cappell KM, Larson B, Sciaky N, Whitehurst AW. 2010. Symplekin specifies mitotic fidelity by supporting microtubule dynamics. Molecular and Cellular Biology 30:5135–5144. doi: 10.1128/MCB.00758-10 Carmena M, Wheelock M, Funabiki H, Earnshaw WC. 2012. The chromosomal passenger complex (CPC): from easy rider to the godfather of mitosis. Nature Reviews. Molecular Cell Biology 13:789–803. doi: 10.1038/ nrm3474 Chaves-Pe´ rez A, Mack B, Maetzel D, Kremling H, Eggert C, Harre´us U, Gires O. 2013. EpCAM regulates cell cycle progression via control of cyclin D1 expression. Oncogene 32:641–650. doi: 10.1038/onc.2012.75 Ciriello G, Miller ML, Aksoy BA, Senbabaoglu Y, Schultz N, Sander C. 2013. Emerging landscape of oncogenic signatures across human cancers. Nature Genetics 45:1127–1133. doi: 10.1038/ng.2762 David CJ, Manley JL. 2010. Alternative pre-mRNA splicing regulation in cancer: pathways and programs unhinged. Genes & Development 24:2343–2364. doi: 10.1101/gad.1973010 Duncan PI, Stojdl DF, Marius RM, Bell JC. 1997. In vivo regulation of alternative pre-mRNA splicing by the Clk1 protein kinase. Molecular and Cellular Biology 17:5996–6001 . doi: 10.1128/mcb.17.10.5996 Edmond V, Moysan E, Khochbin S, Matthias P, Brambilla C, Brambilla E, Gazzeri S, Eymin B. 2011. Acetylation and phosphorylation of SRSF2 control cell fate decision in response to cisplatin. The EMBO Journal 30:510– 523. doi: 10.1038/emboj.2010.333 Fedorov O, Huber K, Eisenreich A, Filippakopoulos P, King O, Bullock AN, Szklarczyk D, Jensen LJ, Fabbro D, Trappe J, Rauch U, Bracher F, Knapp S. 2011. Specific CLK inhibitors from a novel chemotype for regulation of alternative splicing. Chemistry & Biology 18:67–76. doi: 10.1016/j.chembiol.2010.11.009 Grabek KR, Diniz Behn C, Barsh GS, Hesselberth JR, Martin SL. 2015. Enhanced stability and polyadenylation of select mRNAs support rapid thermogenesis in the brown fat of a hibernator. eLife 4. doi: 10.7554/eLife.04517 Gui JF, Lane WS, Fu XD. 1994. A serine kinase regulates intracellular localization of splicing factors in the cell cycle. Nature 369:678–682. doi: 10.1038/369678a0 Harashima H, Dissmeyer N, Schnittger A. 2013. Cell cycle control across the eukaryotic kingdom. Trends in Cell Biology 23:345–356. doi: 10.1016/j.tcb.2013.03.002 Irimia M, Weatheritt RJ, Ellis JD, Parikshak NN, Gonatopoulos-Pournatzis T, Babor M, Quesnel-Vallie`res M, Tapial J, Raj B, O’Hanlon D, Barrios-Rodiles M, Sternberg MJ, Cordes SP, Roth FP, Wrana JL, Geschwind DH, Blencowe BJ. 2014. A highly conserved program of neuronal microexons is misregulated in autistic brains. Cell 159:1511–1523. doi: 10.1016/j.cell.2014.11.035 Jiang K, Patel NA, Watson JE, Apostolatos H, Kleiman E, Hanson O, Hagiwara M, Cooper DR. 2009. Akt2 regulation of Cdc2-like kinases (Clk/Sty), serine/arginine-rich (SR) protein phosphorylation, and insulin-induced alternative splicing of PKCbetaII messenger ribonucleic acid. Endocrinology 150:2087–2097. doi: 10.1210/en. 2008-0818 Karni R, de Stanchina E, Lowe SW, Sinha R, Mu D, Krainer AR. 2007. The gene encoding the splicing factor SF2/ ASF is a proto-oncogene. Nature Structural & Molecular Biology 14:185–193. doi: 10.1038/nsmb1209 Katz Y, Wang ET, Airoldi EM, Burge CB. 2010. Analysis and design of RNA sequencing experiments for identifying isoform regulation. Nature Methods 7:1009–1015. doi: 10.1038/nmeth.1528 Kim Y, Heuser JE, Waterman CM, Cleveland DW. 2008. CENP-E combines a slow, processive motor and a flexible coiled coil to produce an essential motile kinetochore tether. The Journal of Cell Biology 181:411–419. doi: 10.1083/jcb.200802189 Lai MC, Lin RI, Tarn WY. 2003. Differential effects of hyperphosphorylation on splicing factor SRp55. The Biochemical Journal 371:937–945. doi: 10.1042/BJ20021827 Lens SM, Voest EE, Medema RH. 2010. Shared and separate functions of polo-like kinases and aurora kinases in cancer. Nature Reviews. Cancer 10:825–841. doi: 10.1038/nrc2964 Leva V, Giuliano S, Bardoni A, Camerini S, Crescenzi M, Lisa A, Biamonti G, Montecucco A. 2012. Phosphorylation of SRSF1 is modulated by replicational stress. Nucleic Acids Research 40:1106–1117. doi: 10. 1093/nar/gkr837 Long JC, Caceres JF. 2009. The SR protein family of splicing factors: master regulators of gene expression. The Biochemical Journal 417:15–27. doi: 10.1042/BJ20081501 Ly T, Ahmad Y, Shlien A, Soroka D, Mills A, Emanuele MJ, Stratton MR, Lamond AI. 2014. A proteomic chronology of gene expression through the cell cycle in human myeloid leukemia cells. eLife 3:e01630 . doi: 10. 7554/elife.01630

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 17 of 19 Research Article Cell biology Genomics and evolutionary biology

Maslon MM, Heras SR, Bellora N, Eyras E, Ca´ceres JF. 2014. The translational landscape of the splicing factor SRSF1 and its role in mitosis. eLife:e02028. doi: 10.7554/eLife.02028 Matera AG, Wang Z. 2014. A day in the life of the spliceosome. Nature Reviews. Molecular Cell Biology 15:108– 121. doi: 10.1038/nrm3742 Mayr C, Bartel DP. 2009. Widespread shortening of 3’UTRs by alternative cleavage and polyadenylation activates oncogenes in cancer cells. Cell 138:673–684. doi: 10.1016/j.cell.2009.06.016 McDonald WH, Ohi R, Smelkova N, Frendewey D, Gould KL. 1999. Myb-related fission yeast cdc5p is a component of a 40S snRNP-containing complex and is essential for pre-mRNA splicing. Molecular and Cellular Biology 19:5352–5362 . doi: 10.1128/mcb.19.8.5352 Mocciaro A, Rape M. 2012. Emerging regulatory mechanisms in ubiquitin-dependent cell cycle control. Journal of Cell Science 125:255–263. doi: 10.1242/jcs.091199 Moore MJ, Wang Q, Kennedy CJ, Silver PA. 2010. An alternative splicing network links cell-cycle control to apoptosis. Cell 142:625–636. doi: 10.1016/j.cell.2010.07.019 Muraki M, Ohkawara B, Hosoya T, Onogi H, Koizumi J, Koizumi T, Sumi K, Yomoda J, Murray MV, Kimura H, Furuichi K, Shibuya H, Krainer AR, Suzuki M, Hagiwara M. 2004. Manipulation of alternative splicing by a newly developed inhibitor of Clks. The Journal of Biological Chemistry 279:24246–24254. doi: 10.1074/jbc. M314298200 Ninomiya K, Kataoka N, Hagiwara M. 2011. Stress-responsive maturation of Clk1/4 pre-mRNAs promotes phosphorylation of SR splicing factor. The Journal of Cell Biology 195:27–40. doi: 10.1083/jcb.201107093 Oltean S, Bates DO. 2014. Hallmarks of alternative splicing in cancer. Oncogene 33:5311–5318. doi: 10.1038/ onc.2013.533 Pan Q, Shai O, Misquitta C, Zhang W, Saltzman AL, Mohammad N, Babak T, Siu H, Hughes TR, Morris QD, Frey BJ, Blencowe BJ. 2004. Revealing global regulatory features of mammalian alternative splicing using a quantitative microarray platform. Molecular Cell 16:929–941. doi: 10.1016/j.molcel.2004.12.004 Pan Q, Shai O, Lee LJ, Frey BJ, Blencowe BJ. 2008. Deep surveying of alternative splicing complexity in the human transcriptome by high-throughput sequencing. Nature Genetics 40:1413–1415. doi: 10.1038/ng.259 Paronetto MP, Min˜ ana B, Valca´rcel J. 2011. The Ewing sarcoma protein regulates DNA damage-induced alternative splicing. Molecular Cell 43:353–368. doi: 10.1016/j.molcel.2011.05.035 Prasad J, Colwill K, Pawson T, Manley JL. 1999. The protein kinase Clk/Sty directly modulates SR protein activity: both hyper- and hypophosphorylation inhibit splicing. Molecular and Cellular Biology 19:6991–7000 . doi: 10. 1128/mcb.19.10.6991 Satyanarayana A, Kaldis P. 2009. Mammalian cell-cycle regulation: several Cdks, numerous cyclins and diverse compensatory mechanisms. Oncogene 28:2925–2939. doi: 10.1038/onc.2009.170 Shen H, Kan JL, Green MR. 2004. Arginine-serine-rich domains bound at splicing enhancers contact the branchpoint to promote prespliceosome assembly. Molecular Cell 13:367–376 . doi: 10.1016/s1097-2765(04) 00025-5 Shin C, Manley JL. 2002. The SR protein SRp38 represses splicing in M phase cells. Cell 111:407–417 . doi: 10. 1016/s0092-8674(02)01038-3 Staalesen V, Falck J, Geisler S, Bartkova J, Børresen-Dale AL, Lukas J, Lillehaug JR, Bartek J, Lønning PE. 2004. Alternative splicing and mutation status of CHEK2 in stage III breast cancer. Oncogene 23:8535–8544. doi: 10. 1038/sj.onc.1207928 Sterne-Weiler T, Martinez-Nunez RT, Howard JM, Cvitovik I, Katzman S, Tariq MA, Pourmand N, Sanford JR. 2013. Frac-seq reveals isoform-specific recruitment to polyribosomes. Genome Research 23:1615–1623. doi: 10.1101/gr.148585.112 Stumpf CR, Moreno MV, Olshen AB, Taylor BS, Ruggero D. 2013. The translational landscape of the mammalian cell cycle. Molecular Cell 52:574–582. doi: 10.1016/j.molcel.2013.09.018 Tejedor JR, Papasaikas P, Valca´rcel J. 2015. Genome-wide identification of Fas/CD95 alternative splicing regulators reveals links with iron homeostasis. Molecular Cell 57:23–38. doi: 10.1016/j.molcel.2014.10.029 Trapnell C, Williams BA, Pertea G, Mortazavi A, Kwan G, van Baren MJ, Salzberg SL, Wold BJ, Pachter L. 2010. Transcript assembly and quantification by RNA-Seq reveals unannotated transcripts and isoform switching during cell differentiation. Nature Biotechnology 28:511–515. doi: 10.1038/nbt.1621 Tsai YS, Dominguez D, Gomez SM, Wang Z. 2015. Transcriptome-wide identification and study of cancer-specific splicing events across multiple tumors. Oncotarget 6:6825–6839. doi: 10.18632/oncotarget.3145 Vermeulen K, Van Bockstaele DR, Berneman ZN. 2003. The cell cycle: a review of regulation, deregulation and therapeutic targets in cancer. Cell Proliferation 36:131–149 . doi: 10.1046/j.1365-2184.2003.00266.x Wang GS, Cooper TA. 2007. Splicing in disease: disruption of the splicing code and the decoding machinery. Nature Reviews. Genetics 8:749–761. doi: 10.1038/nrg2164 Wang ET, Sandberg R, Luo S, Khrebtukova I, Zhang L, Mayr C, Kingsmore SF, Schroth GP, Burge CB. 2008. Alternative isoform regulation in human tissue transcriptomes. Nature 456:470–476. doi: 10.1038/nature07509 Wang Z, Burge CB. 2008. Splicing regulation: from a parts list of regulatory elements to an integrated splicing code. RNA 14:802–813. doi: 10.1261/rna.876308 Wang K, Singh D, Zeng Z, Coleman SJ, Huang Y, Savich GL, He X, Mieczkowski P, Grimm SA, Perou CM, MacLeod JN, Chiang DY, Prins JF, Liu J. 2010. MapSplice: accurate mapping of RNA-seq reads for splice junction discovery. Nucleic Acids Research 38:e178. doi: 10.1093/nar/gkq622 Wang Y, Chen D, Qian H, Tsai YS, Shao S, Liu Q, Dominguez D, Wang Z. 2014. The splicing factor RBM4 controls apoptosis, proliferation, and migration to suppress tumor progression. Cancer Cell 26:374–389. doi: 10.1016/j. ccr.2014.07.010

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 18 of 19 Research Article Cell biology Genomics and evolutionary biology

Whitfield ML, Sherlock G, Saldanha AJ, Murray JI, Ball CA, Alexander KE, Matese JC, Perou CM, Hurt MM, Brown PO, Botstein D. 2002. Identification of genes periodically expressed in the human cell cycle and their expression in tumors. Molecular Biology of the Cell 13:1977–2000. doi: 10.1091/mbc.02-02-0030. Zhou Z, Fu XD. 2013. Regulation of splicing by SR proteins and SR protein-specific kinases. Chromosoma 122: 191–207. doi: 10.1007/s00412-013-0407-z

Dominguez et al. eLife 2016;5:e10288. DOI: 10.7554/eLife.10288 19 of 19